Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

ARTICLE

pubs.acs.org/EF

Conversion of Anisole Catalyzed by Platinum Supported on


Alumina: The Reaction Network
Ron C. Runnebaum,† Rodrigo J. Lobo-Lapidus,†,§ Tarit Nimmanwudipong,† David E. Block,†,‡ and
Bruce C. Gates*,†

Department of Chemical Engineering and Materials Science, University of California, One Shields Avenue, Davis, California 95616,
United States

Department of Viticulture and Enology, University of California, One Shields Avenue, Davis, California 95616, United States

bS Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV AUTONOMA METROPOLITANA on May 26, 2020 at 20:01:34 (UTC).

ABSTRACT: The conversion of anisole (a compound representative of bio-oils) in the presence of H2 was investigated with a flow
reactor operated at a temperature of 573 K and a pressure of 140 kPa with a platinum on alumina catalyst. Analysis by gas
chromatography mass spectrometry led to the identification of more than 40 reaction products, the most abundant being phenol,
2-methylphenol, benzene, and 2,6-dimethylphenol. The kinetically significant reaction classes were transalkylation, hydrodeoxy-
genation, and hydrogenation. Selectivity-conversion data were used to determine an approximate quantitative reaction network
accounting for phenol, 2-methylphenol, 2-methylanisole, and 4-methylphenol as primary products. Pseudo-first-order rate constants
for the formation of these products are 12, 2.8, 0.14, and 0.039 L/(g of catalyst  h), respectively. A more complete qualitative
network was inferred on the basis of the observed products and the assumption that the reaction classes leading to the most
abundant primary products were responsible for the minor and trace products. The removal of oxygen was evidenced by the
production of benzene.

’ INTRODUCTION prospect of having similar value for biomass conversion. Because


Catalytic conversion of feedstocks formed by pyrolysis of today’s instruments allow analysis of products in much more
cellulosic and lignin-rich biomass is an attractive option for the detail than was reported in early work defining the reactions of
production of renewable liquid transportation fuels.1 3 Although petroleum and coal refining, it is possible to determine reaction
conversion data characterizing complex bio-oil feedstocks have networks that are much more detailed than those reported in
been reported,4 23 there is a lack of fundamental understanding these early investigations. Our goal was to determine such a
of the classes of reactions involved. Sharma et al.6 10 and Vitolo detailed reaction network for a compound representative of
et al.11 investigated the catalytic conversion of various bio-oils, lignin-derived bio-oils and an important functional group
with catalysts such as the zeolite HZSM-5. Adjaye et al.12 14 characterizing compounds in them; consequently, we chose
focused on the conversion of fast pyrolysis bio-oils catalyzed by a compound, anisole, that is aromatic and incorporates an
zeolites and silica aluminas. A few researchers have investigated ether linkage.
the catalytic hydrotreatment of bio-oils,15 23 demonstrating The catalytic hydrotreating of anisole was investigated by
significant removal of oxygen. The complexity of the feedstocks, Bredenberg et al.37 and by Hurff and Klein,38 who used sulfided
however, limits the usefulness of these data for predicting hydrotreating catalysts; reaction networks, however, were either
the conversions of other feedstocks. In these investigations, the not reported or limited to few products. Adjaye and Bakhshi,12,14,34
individual products found in the conversion of bio-oils have Hurff and Klein,38 Samolada et al.,17 Sharma and Bakhshi,6 10 and
sometimes been grouped into classes (lumps), and approximate Elliott et al.15 reported mass balance closures, but others have not.
reaction networks showing conversion to the lumps have been The following paper is part of a series39 42 concerned with
reported. various bio-oil compounds; early results for anisole have been
The conversion of individual, prototypical bio-oil compounds communicated.39 We now report details, which include mass
has also been investigated by a number of researchers. The balances, selectivity-conversion plots with error bounds, rate
catalytic reaction networks presented by these researchers, which constants of product formation, and a quantitative reaction net-
were determined from conversions measured in the absence of work, characterizing the conversion of anisole catalyzed by a
H2, are limited either to primary products (or occasionally supported metal, Pt/Al2O3, in the presence of H2. We also report
primary and secondary products) or to products that have been the design of a reactor system that includes the capability for
lumped into classes of similar compounds (e.g., phenol, alkyl- extensive online analysis of gas products by gas chromatography
phenol, and alkylated ether lumps).24 35
Investigations of the reactions of individual compounds have Received: July 21, 2011
been of value in advancing the understanding of petroleum Revised: August 25, 2011
and coal-liquid hydroprocessing reactions,36 and they offer the Published: August 28, 2011

r 2011 American Chemical Society 4776 dx.doi.org/10.1021/ef2010699 | Energy Fuels 2011, 25, 4776–4785
Energy & Fuels ARTICLE

Figure 1. Schematic representation of once-through, packed-bed flow reactor system. Abbreviations: TC, thermocouple; PI, pressure indicator; PT,
pressure transducer; PG, analogue pressure gauge; BPR, back-pressure regulator; MFM, mass flow meter; MFC, mass flow controller.

and at-line analysis of liquid products by gas chromatography were used to measure and control gas flow rates. N2 was used as the
with identification of products by gas chromatography/mass carrier gas. H2, when used as a reactant, was mixed into the carrier gas
spectrometry; we have used this system to evaluate reaction stream via a Swagelok union cross (4-way) downstream of the mass flow
products and establish the reaction network. controllers. The gas stream was mixed with vaporized anisole reactant at
the inlet of the reactor in the top zone of a three-zone furnace. At the top
’ EXPERIMENTAL METHODS of the downflow reactor, a Swagelok union cross provided three inlets to
the reactor, one for gases, one for liquid fed through a capillary tube
Materials. Anisole (methoxybenzene, 99.8%) was purchased from (stainless steel, O.D. = 1/16 in., I.D. = 0.020 in., Model U-104, Upchurch
Sigma-Aldrich. Pt/γ-Al2O3 powder (<100 mesh) containing 1 wt % Pt Scientific), and one for a thermocouple (Model HKMTSS-040U-12,
was acquired from Sigma-Aldrich. N2 (99.997%, from Praxair), further Omega Engineering), which was placed near the tip of the capillary tube
purified by passage through a BHT-4 hydrocarbon trap (Agilent to measure the temperature of vaporization of the liquid feed. From
Technologies) to remove traces of hydrocarbon, was used as an inert the outlet of the reactor, the product stream flowed through a Swagelok
carrier gas in the gas solid reaction system and in the gas chromato- (3-way) union tee, which also allowed mounting of a thermocouple
graphs. H2 (99.999%), generated from deionized water by electrolysis in (Model KQSS-116U-18, Omega Engineering) inside the reactor im-
a model 40H gas generator (Domnick Hunter), was used as a reactant in mediately downstream of the packed bed.
the catalytic conversion experiments and in the gas chromatographs. The reactor was mounted in a three-zone furnace (Series 3210,
Helium (99.995%, Praxair) was purified by passage through a BHT-4 Applied Test Systems), which was designed for precise temperature
trap (Agilent Technologies) to remove traces of hydrocarbon, a BMT-4 control. Five thermocouples were used to measure temperatures inside
trap (Agilent Technologies) to remove moisture, and a BOT-4 trap (at the two locations mentioned above) and outside the reactor. The
(Agilent Technologies) to remove oxygen prior to use in the gas temperatures of the vaporization zone and packed bed were controlled
chromatographs. Zero-air (<0.1 ppm hydrocarbon (as methane)) was by using the measurements of temperature determined with thermo-
generated from house air in a model UHP-10ZAW zero air generator couples in the reactor. Thermocouple wells extending axially into each zone
(Domnick Hunter) and was used in the gas chromatograph flame of the furnace enabled positioning of thermocouples near the outer wall
ionization detectors. The BET surface area of the catalyst was measured of the reactor and monitoring of temperatures in each zone at the outer
with a Micromeritics ASAP 2020 instrument, and the dispersion of the reactor wall. The temperature of the zone downstream of the packed bed
platinum in the catalyst was measured with a Micromeritics Autochem II was controlled by using the signal of the thermocouple mounted at the
Chemisorption Analyzer. outside wall of the reactor. The temperatures of the three heating zones
Flow Reactor System. Catalytic reaction experiments were carried were controlled separately.
out with a once-through packed-bed flow reactor (316 L stainless steel, The packed bed, holding 4.1 401.5 mg of catalyst particles, was
O.D. = 0.5 in., wall thickness = 0.035 in.). A schematic diagram of the typically 10 to 30 mm long. Glass wool and the packed bed of catalyst
flow system is shown in Figure 1. particles were positioned at the midpoint of the reactor and furnace. To
The liquid reactant was preloaded at room temperature into a syringe provide nearly flat radial velocity and temperature profiles in the bed, the
pump (Model 500D, Teledyne-ISCO), which was used as a liquid feed catalyst particles were mixed with particles of inert α-Al2O3 (A634-3
reservoir. Mass flow controllers (Model SLA5850S, Brooks Instruments) alundum, Fisher, 90 mesh particle size).

4777 dx.doi.org/10.1021/ef2010699 |Energy Fuels 2011, 25, 4776–4785


Energy & Fuels ARTICLE

After exiting the reactor, the product stream flowed through stainless SI Table 1) in the gas calibration standard, were used to quantify
steel tubing that was heated by electrical heating tape (HTC-120, products in the gas stream.
Omega), extending from the exit of the reactor to the inlet of a Catalyst Loading, Preparation, and Testing. The Pt/Al2O3
condenser. Water at approximately 10 °C was used as a coolant, and catalyst was held in place in the reactor with approximately 0.1 g of
the outlet gas stream exited the condenser at approximately 12 °C, quartz wool mounted near the axial center of the reactor. A bed of
as measured by a thermocouple (KQMSS-020G-6, Omega). Liquid approximately 2 g of α-Al2O3 was placed on the quartz wool; down-
samples were collected at intervals of approximately 20 min from the stream of it was a bed of catalyst particles mixed with 1 2 g of α-Al2O3,
bottom of the condenser and weighed. The flow rate of the gas product and upstream of that was another bed of 2 g of α-Al2O3 particles. The
stream was monitored with a SLA 7860S mass flow meter (Brooks catalyst was heated in N2 flowing at 70 mL/min at NTP and H2 (flowing
Instruments). at a rate of 30 mL/min at NTP) at a rate of 5 °C/min to 573 K and held
Safety features enabled redundant monitoring of key control points. at this temperature for 30 min prior to the start of reactant flow. The
For instance, both pressure transducers (PTI-S-NG250-22AQ, Swagelok) catalyst mass was in the range of 4.1 401.5 mg.
and analogue pressure gauges (PGI-63B-PG200-LAQX-J, Swagelok) The catalytic conversion of anisole was carried out at 573 K and
were placed upstream of the reactor and downstream of the condenser; approximately 140 kPa; the anisole flow rate was 0.03 mL/min, and the
the pressure in the system was controlled by a manual back-pressure N2 and H2 flow rates were 70 and 30 mL(NTP)/min, respectively. The
regulator (44-2300 Series, Tescom). Reactor temperatures were mea- range of the weight hourly space velocity (WHSV) was 17.6 437 (g of
sured in each zone, both inside and outside of the reactor; upper limits anisole)/(g of catalyst  h), varied by changing the catalyst mass. Fresh
were included in the control program for automatic shut-off of power to catalyst was used for each experiment. Experiments were typically run for
the furnace. 6 h time-on-stream.
Product Analysis. Analysis of the products in the liquid samples was Data Analysis. Conversions of anisole were determined on the
performed with an Agilent 7890A gas chromatograph (GC) equipped basis of mass balance calculations for the reaction system. The terms in
with a 5975C mass selective detector (MSD) incorporating a quadru- the mass balance included the flow rate of anisole into the reactor and
pole mass analyzer. Analyte separation was achieved with a DB-624 the flow rates of the liquid condensate and gas streams out of the
Agilent J&W column (0.25 mm/60 m/1.4 μm) with 2.4 mL/min helium condenser; the mass of the packed bed was measured before and after
carrier gas flow rate and a temperature ramp program. Peaks were each experiment to allow a determination of the accumulation of mass in
identified by using the Agilent quality match factor (Qual), which is based it. The mass of reactant entering the reactor was calculated from the
on how well the ion ratios match those of the NIST EI mass spectral volumetric flow rate and density of the reactant (flowing at approxi-
library database standard; each compound except water and methanol mately NTP). The mass of liquid condensate was determined periodi-
scored greater than 90. Retention times of some of the compounds were cally by measuring the mass of liquid collected in each vial and summing
confirmed by use of calibration samples. Product quantification was over the samples. The mass of the reactant leaving the condenser in the
carried out with an Agilent 7890A GC using a flame ionization detector gas phase was calculated from the flow rate of the gas and by using an
(FID); analyte separation was achieved with a DB-624 Agilent J&W Antoine vapor pressure correlation and assuming that the inert gas was
column (0.25 mm/60 m/1.4 μm). The operating parameters of the GC- saturated with the reactant at the exit temperature of the condenser.
FID were developed to obtain essentially the same peak retention times Because the condenser did not drain perfectly, the mass of liquid
and separations as with the GC-MSD. The FID was operated at 300 mL/ remaining in the condenser at the end of each experiment was
min zero-air and 25 mL/min H2 flow rates. Calibration curves, each with determined from the mass of liquid fed into the system prior to the
at least five points, were used to quantify the following products: collection of liquid at the bottom of the condenser.
benzene, cyclohexanone, phenol, 2-methylanisole, 4-methylanisole, Conversions (X) were calculated as the molar flow rates of anisole
2-methylphenol, 4-methylphenol, and 2,6-dimethylphenol. Other pro- reactant consumed divided by the total molar flow rate of anisole fed.
ducts were analyzed semiquantitatively on the basis of results obtained The mass of products, taken to be the mass of reactant consumed, was
for chemically similar components (containing the same numbers of determined by quantifying GC-FID peak areas. The number of moles of
carbon and oxygen atoms). reactant consumed in any experiment was calculated from the mass of
A GC (Agilent 7890A), equipped with three sample loops, five reactant consumed and the molecular weight of the reactant. Selectivity is
columns, and three detectors (one FID and two TC detectors), was defined here as the molar flow rate of product divided by molar flow rate of
used to analyze the gas effluent from the condenser. Each of the sample reactant consumed. Yield is defined here as conversion times selectivity.
loops was 500 μL in volume. Hydrocarbons were separated in an HP-Al/ Initial conversion data were obtained by extrapolating the conversion
S (19095P-S25, 50 m long/0.53 mm in I.D./15 μm in film thickness) vs time-on-stream data to zero time on stream by using a linear fit. These
Agilent J&W capillary column with a helium carrier gas flow rate of initial conversion data were used to find the fraction of anisole reactant
27 mL/min, and the components were quantified by analysis with an unconverted as a function of inverse WHSV, as well as yield as a function
FID. The FID was operated at 300 mL/min zero-air and 25 mL/min H2 of conversion and selectivity as a function of conversion.
flow rates. Fixed gases and methane were separated with three Agilent Identification of Primary and Nonprimary Products. Selectivity vs.
1/8-in stainless-steel packed columns mounted in series: HayeSep Q (6 ft), conversion plots were used to determine which products were primary
HayeSep N (8 ft), and Molsieve 13X (10 ft), and the components were and which were nonprimary, on the basis of the ordinate-intercepts of
quantified with a TCD with helium carrier gas at a pressure of 70 kPa. H2 the linear regression line through the data points; data that determine an
was separated in a 4-ft Molsieve 13X packed column and quantified with ordinate axis intercept of the regression line significantly different from
a TCD using N2 carrier gas. Temperature and carrier gas flow ramps zero (with 95% confidence limits) are considered to be indicative of
were used to improve the retention time and separation of compounds. primary products.43,44
Gas samples were taken online at approximately 30 min intervals; each of Determination of Reaction Rates from Differential Conversion Data.
the sample loops was actuated, nearly simultaneously, to load the flow Data were plotted to demonstrate that some conversions were differential.
paths to each of the detectors and to characterize the gas stream. Peak The pseudo-first-order rate constant for the overall disappearance of anisole
identification was carried out by comparison with a NIST traceable was determined from a semilogarithmic plot of fraction of reactant
Agilent Technologies/Praxair RGA gas calibration standard (5184- unconverted, (1 X), as a function of inverse WHSV. The pseudo-first-
3545, Lot No. 091509R). Single-point calibrations, determined using order rate constants for the production of primary products were deter-
data collected with compounds (listed in the Supporting Information, mined from plots of conversion vs inverse space velocity.

4778 dx.doi.org/10.1021/ef2010699 |Energy Fuels 2011, 25, 4776–4785


Energy & Fuels ARTICLE

Table 1. Overall Mass Balance Closure Dataa


experiment mass balance time on mass of liquid mass of liquid collected from mass of reactant leaving mass of liquid remaining in
number closure (%) stream (min) fedb (g) condenserb (g) condenser as vaporb (g) condenser (g), estimated

1 94.7 347 10.4 9.04 0.26 0.5


2 99.5 286 8.56 7.62 0.20 0.5
3 99.8 403 12.0 11.2 0.34 0.5
4 98.9 382 11.4 10.5 0.32 0.5
5 99.2 358 10.7 9.81 0.30 0.5
6 100.5 381 11.4 10.7 0.32 0.5
7 101.0 170 5.06 4.47 0.14 0.5
a
The data in each row represent results from an individual experiment. b The data showing liquid fed, liquid collected from condenser, and reactant
leaving the condenser in the vapor phase are cumulative values for the length of the experiment.

’ RESULTS space velocities is shown in Figure 2. These data demonstrate


catalyst deactivation. In a typical experiment, the deactivation
Catalyst Physical Properties. The BET surface area of the led to a decline in conversion of about 50% in 6 h. The yields
catalyst was found to be 206 ( 1 m2/g, and the dispersion of the of various products vs time on stream do not suggest that
platinum was found to be approximately 0.25. deactivation affects a particular reaction class more strongly
Mass Balance Closures. Mass balance closures were typically than the others. Initial conversions were determined by extra-
99.5 ( 1.0% (and they were generally greater than 95%) in polating the data to zero time on stream. Error bars on the
experiments with anisole conversions up to 17% (Table 1). The data are shown for experiments carried out with a WHSV of
reactant and most products have low vapor pressures at the exit 34 (g of reactant)/(g of catalyst  h). The data indicate that
temperature of the condenser, and we, therefore, infer that they higher space velocities lead to lower initial conversions, as
were almost completely condensed. The mass of reactant leaving expected.
the condenser in the vapor phase was determined to be between Development of a Quantitative Reaction Network. Plots of
2.5 and 2.8% of the mass fed into the reactor, and it is accounted selectivity (molar flow rate of product divided by molar flow rate
for in the mass balance calculation (Table 1). The contributions of reactant consumed) as a function of conversion were used to
of the products exiting the condenser in the vapor phase, which identify which products were primary and which were not.43,44
generally have lower vapor pressures at the temperature of the Data for each product were fitted with a straight line and
condenser, constitute a negligible contribution to the overall extrapolated to zero conversion; intercepts of linear regression
mass balance. As a result, mass balance closures were not signi- lines significantly different from zero selectivity at zero conver-
ficantly different from 100% even when a quantitative accounting sion (determined with 95% confidence limits) indicate primary
for products exiting the condenser in the gas phase was omitted. products. The selectivity-conversion data observed for anisole
The changes in mass of the packed bed were negligible relative to (Figure 3) indicate that phenol, 2-methylphenol, 4-methyl-
other terms in the mass balance and did not contribute sig- phenol, 2-methylanisole, and 2,6-dimethylphenol were pri-
nificantly to the overall mass balance. mary products and that benzene and cyclohexanone were
Major, Minor, and Trace Products of Anisole Conversion. higher-order products. We infer, however, that 2,6-dimethyl
Dozens of products were identified in the conversion of anisole phenol was produced from anisole by a sequence of reactions
catalyzed by Pt/Al2O3. Data were obtained at conversions as low and, in small quantities, from feed impurities (<0.2%) such as
as 0.01 to determine primary reactions and the associated kinetics methylphenols.
parameters. Product formation data were also obtained at higher The first step in the determination of a reaction network is to
conversions, up to X = 0.2, to identify nonprimary products. As a infer the likely reaction classes from the identities of the products.
control, experiments were performed using a bed of only α- Thus, analysis of the products can give an indication of the reaction
Al2O3 (without catalyst); essentially no conversion (X < 0.001) classes in the conversion of anisole catalyzed by Pt/Al2O3.
was observed under these conditions. The most abundant products point to the following important
The most abundant compounds identified at an anisole (i.e., kinetically significant) reaction classes: hydrogenolysis (indi-
conversion of approximately 0.2 are listed in Table 2. The major cated, for example, by the formation of phenol and methane),
products include methane, benzene, cyclohexanone, 2-methyl- hydrodeoxygenation (indicated, for example, by the formation of
phenol, 2-methylanisole, phenol, and 2,6-dimethylphenol. The benzene), hydrogenation (indicated, for example, by the formation
conversions to these products were determined quantitatively. of cyclohexanone), and methyl group transfer (transalkylation)
Other products formed in relatively high yields include 4-methyl- (indicated, for example, by the formation of 2-methylphenol).
phenol, methoxycyclohexane, and 4-methylanisole. Water and We infer that phenol is formed in both hydrogenolysis
methanol were products observed by GC-MS but not quantified. and transalkylation reactions; other major products, including
Trace products (identified only qualitatively) included methyl- 2-methylanisole and 2-methylphenol, likely result from single
cyclohexanone, cyclohexane, methylcyclohexane, 1,1-dimethyl- methyl group transfer reactions. 2,6-Dimethylphenol is inferred
cyclohexane, o-xylene, p-xylene, 2,6-dimethylanisole, 2,4-dimethyl- to have been formed in sequential transalkylation reactions.
phenol, and trimethylphenol. Methanol and benzene are inferred to have been formed by
Conversion of Anisole as a Function of Time on Stream hydrogenolysis of the C O bond, referred to here as hydro-
and of Inverse Space Velocity. The conversion of anisole deoxygenation. Cyclohexanone is inferred to have been formed
catalyzed by Pt/Al2O3 as a function of time on stream at various by a sequence of reactions: removal of the methyl group bonded
4779 dx.doi.org/10.1021/ef2010699 |Energy Fuels 2011, 25, 4776–4785
Energy & Fuels ARTICLE

Table 2. Most Abundant Products and Trace Products Formed in the Conversiona of Anisole Catalyzed by Pt/Al2O3
(Compounds Identified in Liquid Product)b
classification based on
product abundance in product stream basis for identification of product basis for quantification of product

benzene major comparison of GC trace with that of standard sample FID calibration curve
methoxycyclohexane minor MS EI database n/a
cyclohexanone major comparison of GC trace with that of standard sample FID calibration curve
2-methylanisole major comparison of GC trace with that of standard sample FID calibration curve
4-methylanisole minor comparison of GC trace with that of standard sample n/a
phenol major comparison of GC trace with that of standard sample FID calibration curve
2-methylphenol major comparison of GC trace with that of standard sample FID calibration curve
4-methylphenol minor comparison of GC trace with that of standard sample n/a
2,6-dimethylphenol major comparison of GC trace with that of standard sample FID calibration curve
cyclohexane trace comparison of GC trace with that of standard sample n/a
methylcyclohexane trace MS EI database n/a
1,1 dimethylcyclohexane trace MS EI database n/a
p-xylene trace MS EI database n/a
o-xylene trace MS EI database n/a
2-methylcyclohexanone trace MS EI database n/a
2,6-dimethylanisole trace MS EI database n/a
2,4-dimethylphenol trace MS EI database n/a
trimethylphenol trace MS EI database n/a
water n/a MS EI database n/a
methanolc n/a MS EI database n/a
a
Conversion conditions: temperature, 573 K; WHSV, 4.46 (g of reactant)/(g of catalyst  h); initial conversion of anisole was approximately 0.2. The
retention times of the compounds identified in the liquid samples were consistent with the retention times of the compounds determined in experiments
with authentic calibration samples. b Compounds were identified with the NIST EI mass spectral database and by requiring that the Agilent quality
match factor (Qual) score be greater than 90; other isomers, for example, other dimethylanisole isomers, are possible. Some compounds were confirmed
by comparison with a GC trace characterizing standard samples. c Methanol was observed in liquid samples; its concentration in the gas phase was not
quantified.

The trace products are inferred to have been formed by


sequential reactions. These products can be accounted for by
the following reaction classes: transalkylation and hydrogenation
of the aromatic ring, leading to methylcyclohexanone, and hydro-
deoxygenation, hydrogenation, and transalkylation, accounting
for cyclohexane and methylcyclohexane. Multiple transalkylation
reactions account for methylphenols and methylanisoles with
multiple methyl substituents (e.g., various dimethylphenol iso-
mers and dimethylanisole isomers).
A partial statement of the reaction network, representing just
the primary products and other important products, is shown in
Figure 4. According to this representation, hydrogenolysis,
methyl group transfer, hydrodeoxygenation, and hydrogenation
are the kinetically significant reaction classes in the anisole con-
version catalyzed by Pt/Al2O3 in the presence of H2. Specifically,
methyl group transfer to methylphenols produced dimethylphe-
Figure 2. Change in anisole conversion during operation of a flow reactor nols, which we infer eventually led to trimethylphenols and
at 573 K and at various values of WHSV (WHSV in units of (g of reactant)/ tetramethylphenols. Methylbenzene (toluene) could be pro-
(g of catalyst  h)): 18 (b), 34 ()), 150 (O), and 440 (9). An average of duced from either methylphenol or methylanisole by hydro-
duplicate values is shown for WHSV = 34 (g of reactant)/(g of catalyst  h). deoxygenation reactions, which also lead to the formation of
benzene from phenol or anisole.
to oxygen by hydrogenolysis or dealkylation and hydrogenation The data allow a quantitative characterization of the overall
of the aromatic ring. conversion of anisole catalyzed by Pt/Al2O3. The overall anisole
Transalkylation reactions are inferred to have produced the conversion is well represented by first-order kinetics (Figure 5).
minor products 4-methylphenol and 4-methylanisole. Sequential The pseudo-first-order rate constant for the disappearance of
hydrogenation reactions of the aromatic ring are inferred to have anisole was found to be 19 L/(g of catalyst  h) under the
produced methoxycyclohexane. conditions stated in the caption of Figure 2. Using estimates of
4780 dx.doi.org/10.1021/ef2010699 |Energy Fuels 2011, 25, 4776–4785
Energy & Fuels ARTICLE

Figure 3. Selectivity for the formation of several products in the conversion of anisole catalyzed by Pt/Al2O3 in the presence of H2 at 573 K; error bars
were determined in replicate experiments at conversions of approximately 0.07 and 0.16. Data for each product were fitted with a straight line and
extrapolated to zero conversion; intercepts of regression lines significantly different from zero selectivity at zero conversion (analyzed with 95%
confidence limits) indicate primary products. Primary products in this case are phenol (a), 2-methylphenol (b), 2-methylanisole (f), and 2,6-
dimethylphenol (d), and those with intercepts not significantly different from zero (determined with 95% confidence limits) are considered to be
evidence of higher-order products, in this case benzene (c) and cyclohexanone (e). The selectivity-conversion plot for 4-methylphenol (ESI Figure 1)
indicates that it is a primary product.

the diameter and density of catalyst particles, the average dia- The product that formed fastest from anisole was phenol. This
meter of the catalyst pores, and the reactant diffusivity, we judged product was formed more than four times faster than 2-methyl-
on the basis of the Weisz-Prater criterion, assuming spherical phenol, and the other products were formed even more slowly.
particles and first-order kinetics, that influence of intraparticle The formation of 2-methylanisole was an order of magnitude
mass transport was negligible. slower than the formation of 2-methylphenol. The formation of
Data characterizing the formation of some of the products are 4-methylphenol was nearly 2 orders of magnitude slower than
also well represented by first-order kinetics. The results (Figure 6) the formation of 2-methylphenol, which indicates a kinetic
allow the partially quantified statement of the reaction network preference for the substitution at the ortho position. Each of
shown in Figure 4, which includes the pseudo-first-order rate the products is possibly formed by a single reaction; however,
constants for each of the reactions, for example, the methyl group these products could also be formed by sequences of reactions.
transfer that results in the formation of 2-methylphenol from In contrast, we infer that 2,6-dimethylphenol was formed only in
anisole. The pseudo-first-order rate constants for the formation a sequence of methyl-group transfers. The primary product, for
of these products are summarized in Figure 4. instance, can be inferred to be 2-methylphenol or 2-methylanisole.
4781 dx.doi.org/10.1021/ef2010699 |Energy Fuels 2011, 25, 4776–4785
Energy & Fuels ARTICLE

Figure 4. Reaction network to primary products and to other important products based on analysis of selectivity-conversion plots for the conversion of
anisole and H2 catalyzed by Pt/Al2O3 at 573 K. H2, as a reactant, is omitted for simplicity. 2,6-Dimethylphenol appears (on the basis of the selectivity-
conversion plot) to be a primary product, but we infer that it must have been formed by a sequence of reactions. Pseudo-first-order rate constants for
the individual primary products formed in the conversion of anisole catalyzed by Pt/Al2O3 are in units of L/(g of catalyst  h). The reaction temperature
was 573 K. These rate constants were determined from the conversions to the individual products, which, for example, in the case of phenol, may be
formed in more than one reaction pathway, as shown here.

determined to be k = 0.86 L/(g of catalyst  h), and that found


for cyclohexanone formation was k = 0.53 L/(g of catalyst  h).
Presuming that methyl group transfer, hydrodeoxygenation,
hydrogenolysis, and hydrogenation are the important reaction
classes and by recognizing which compounds were primary
products and which were not, we inferred the more extensive
reaction network of Figure 8. As this network shows, most of the
products in the conversion of anisole were likely formed in
sequential reactions; they were not primary products. However, the
data are not sufficient to fully elucidate the nonprimary products.
The data are not sufficient to identify all of the primary
reactions. For example, phenol could have been formed as a
primary product via several pathways, each with a different
coproduct: methane, 2-methylanisole, and 4-methylanisole. Be-
cause the selectivity for formation of neither 2-methylanisole nor
4-methylanisole matches that of phenol, we infer that the primary
pathway for phenol formation is hydrogenolysis of anisole (to
Figure 5. Demonstration of first-order kinetics of overall conversion give phenol and methane).
of anisole catalyzed by Pt/Al2O3 under the conditions stated in the text. 2-Methylphenol, as well as 4-methylphenol, can also be formed
Error bars of replicates are shown at inverse WHSV = 0.056 and 0.029 via multiple pathways: intramolecular transalkylation of anisole
(g of catalyst  h)/(g of anisole).
or intermolecular transalkylation involving phenol and anisole.
The further addition of methyl substituents to the aromatic ring
The formation of benzene and of cyclohexanone, although to produce dimethylphenols, trimethylphenols, and so forth
they are evidently not primary products, is in each case adequately allows for a number of possible pathways. 2-Methylanisole and
represented by first-order kinetics (Figure 7). The pseudo- 4-methylanisole are inferred to be formed primarily from single
first-order rate constant characterizing benzene formation was methyl group transfer reactions.
4782 dx.doi.org/10.1021/ef2010699 |Energy Fuels 2011, 25, 4776–4785
Energy & Fuels ARTICLE

Figure 6. Semilogarithmic plots of the conversion of anisole in the presence of H2 to give phenol (O), 2-methylphenol (b), 2,6-dimethylphenol (0),
2-methylanisole (9), and 4-methylphenol (1) at 573 K. The term (1 Xi) represents the conversion to product i. Conversion was varied by changing
the catalyst mass in the packed bed.

reactor ranging from less than 0.05 to 1.0 h. For comparison, our
residence times ranged from 2.3  10 3 to 56.7  10 3 h.
In their experiments with acidic catalysts, Chantal et al.32 and
Zhu et al.35 also observed the removal of oxygen from anisole and
the formation of aromatic products. The highest selectivity to
aromatics indicated by their data was 0.05,32 which is lower than
our observed selectivity to benzene of almost 0.10. Chantal
et al.32 and Zhu et al.35 used higher temperatures and higher
average residence times than ours, and we suggest that the
differences in severity between their conditions and ours account
in part for the different product distributions, but we attribute the
differences primarily to the presence of the noble metal in our
catalyst and the lack of such a metal in theirs.
In experiments with a bifunctional hydrotreating catalyst,
sulfided Ni Mo/SiO2 Al2O3, which contains an acidic sup-
Figure 7. Semilogarithmic plots of the conversion of anisole in the
port, Bredenberg et al.37 also observed the formation of methyl-
presence of H2 to give cyclohexanone (O) and benzene (b) at 573 K.
The term (1 Xi) represents the conversion of anisole to product i. phenols and methylanisole in the conversion of anisole at 573 K.
Conversion was varied by changing the catalyst mass in the packed bed. They also reported a selectivity for benzene formation of 0.055 at
an anisole conversion of 0.76. The data cannot be directly
Benzene was also possibly formed via parallel pathways, both compared with ours, which are characterized by a maximum
involving hydrodeoxygenation. From the (qualitative) observa- observed selectivity for benzene formation of 0.10, because the
tion of methanol, we infer that hydrodeoxygenation of anisole conversions are dramatically different. The conversion and yield
produces methanol and benzene. Hydrodeoxygenation of phe- data reported by Hurff and Klein38 indicate a selectivity for
nol can also produce benzene and water. Hydrogenation of benzene formation at high conversion similar to that of Breden-
phenol has been shown to produce cyclohexanone.45,46 These berg et al.,37 although a precise value cannot be estimated at low
reactions with hydrogen likely occurred with methyl-substituted conversions because of the scale of the plot in the publication.38
phenols and anisoles to form methyl-substituted benzenes. Thus, the routes by which oxygen was removed from anisole in
In summary, the reaction network of Figure 8 accounts for our the reported acid-catalyzed reactions are not clear from the
data and accounts for the kinetically important reaction classes, information presented by the authors. We emphasize that in
but we recognize that it is an approximation that could be our case we attribute the oxygen removal largely to the role of
improved with the availability of more extensive data. platinum in activating H2 for hydrodeoxygenation reactions.
There are numerous examples of similar reactions catalyzed by
noble metals, exemplified the hydrodeoxygenation of guaiacol
’ DISCUSSION catalyzed by rhodium, palladium, and platinum supported on
The catalytic conversion of anisole catalyzed by platinum zirconia.47 Ours is evidently the first report of the hydrode-
supported on γ-Al2O3, an acidic support, shows some similarity oxygenation of anisole catalyzed by a supported noble metal.
to anisole conversion catalyzed by solid acids without noble Hydroprocessing catalysts incorporating metals such as mo-
metals. For example, Chantal et al.,32 using HZSM-5 as a catalyst, lybdenum have also been used for hydrodeoxygenation. For
observed high yields of phenol, methylphenols, and dimethyl- example, Hurff and Klein38 and Bredenberg et al.37 used sulfided
phenols at 623 K with a residence time in a plug-flow reactor of CoMo/γ-Al2O3 for removal of oxygen from anisole, observing
0.83 h. Using HZSM-5 as a catalyst, Zhu et al.35 observed the benzene as a major product at high conversions in a batch
formation of phenol, methylphenols, and methylanisoles in high reactor; the benzene was evidently formed from phenol, which
yields at 673 K and over a range of residence times in a plug-flow was a major primary product according to their reaction network.
4783 dx.doi.org/10.1021/ef2010699 |Energy Fuels 2011, 25, 4776–4785
Energy & Fuels ARTICLE

Figure 8. Reaction network for the conversion of anisole and H2 catalyzed by Pt/Al2O3 at 573 K. Hydrodeoxygenation, hydrogenolysis, and
hydrogenation reactions are represented by dashed arrows, and methyl group transfer reactions are represented by solid arrows.

In contrast, our results show the catalytic formation of benzene at Although the reaction network could be elucidated in more
relatively low conversions of anisole, consistent with the catalytic detail than we have presented, with the benefit of additional data
role of the platinum. (such as could be obtained with intermediates in the feed
stream), an important practical inference that is already clear
on the basis of our results is that selective oxygen removal from
’ CONCLUSIONS compounds such as anisole may require both (a) catalysts with
In comparison with the work of earlier investigators who functions that are active for hydrodeoxygenation (such as noble
considered catalytic anisole conversion, we have been able to metals) and (b) substantially higher H2 partial pressures than we
develop a more extensive reaction network, including quanti- have investigated in this work.
fication of rates of formation of primary products. The data
enable us to infer relative contributions of the various reaction ’ ASSOCIATED CONTENT
classes. We infer the platinum catalyzed hydrogenolysis of the
anisole to produce methane and phenol, which is a kinetically bS Supporting Information. Table containing the composi-
significant reaction under our conditions. All other primary tion of the gas calibration standard and a figure showing 4-methyl-
products were apparently produced by methyl group transfer phenol selectivity-conversion data. This material is available free of
reactions. Higher-order products that were observed in rela- charge via the Internet at http://pubs.acs.org.
tively high yields include benzene, formed by hydrodeoxygena-
tion of anisole or phenol, and cyclohexanone, formed by ’ AUTHOR INFORMATION
hydrogenation of phenol. We infer that reactions in both of Corresponding Author
these classes were catalyzed by the platinum. Methyl group *E-mail: bcgates@ucdavis.edu.
transfer apparently catalyzed the formation of numerous higher-
order products such as dimethylanisoles, trimethylphenols, and Present Addresses
§
pentamethylbenzenes, formed, for example, from methylanisole Argonne National Laboratory, Chemical Sciences and Engi-
and anisole, dimethylphenol and anisole, and tetramethylbenzene neering Division, 9700 South Cass Ave, CSE/205, Chicago, IL
and anisole, respectively. 60439, USA.
4784 dx.doi.org/10.1021/ef2010699 |Energy Fuels 2011, 25, 4776–4785
Energy & Fuels ARTICLE

’ ACKNOWLEDGMENT (31) Kallury, R. K. M. R.; Restivo, W. M.; Tidwell, T. T.; Boocock,


We thank Jennifer Heelan of the University of California, D. G. B.; Crimi, A.; Douglas, J. J. Catal. 1985, 96, 535–543.
(32) Chantal, P. D.; Kaliaguine, S.; Grandmaison, J. L. Appl. Catal.
Davis, for her help with the analytical instrumentation and 1985, 18, 133–145.
Professor Alexander Katz of the University of California, Berkeley, (33) Kallury, R. K. M.; Tidwell, T. T.; Boocock, D. G. B.; Chow,
for providing access to instruments for determination of the D. H. L. Can. J. Chem. 1984, 62, 2540–2545.
catalyst physical properties. Support of this work was provided by (34) Adjaye, J. D.; Bakhshi, N. N. Biomass Bioenergy 1995, 8,
Chevron Technology Ventures, a division of Chevron U.S.A., 131–149.
Inc. An Agilent Technologies Foundation Research Project Gift (35) Zhu, X.; Mallinson, R. G.; Resasco, D. E. Appl. Catal., A: Gen.
provided a GC7890 Refinery Gas Analyzer. T.N. thanks the 2010, 379, 172–181.
Fulbright Foundation for an Open Competition scholarship. (36) Girgis, M. J.; Gates, B. C. Ind. Eng. Chem. Res. 1991, 30,
2021–2058.
(37) Bredenberg, J. B.; Huuska, M.; R€aty, J.; Korpio, M. J. Catal.
’ REFERENCES 1982, 77, 242–247.
(1) Bridgwater, A. V.; Cottam, M. L. Energy Fuels 1992, 6, 113–120. (38) Hurff, S. J.; Klein, M. T. Ind. Eng. Chem. Res. 1983, 22, 426–430.
(2) Huber, G. W.; Iborra, S.; Corma, A. Chem. Rev. 2006, 106, (39) Runnebaum, R. C.; Nimmanwudipong, T.; Block, D. E.; Gates,
4044–4098. B. C. Catal. Lett. 2011, 141, 817–820.
(3) St€ocker, M. Angew. Chem., Int. Ed. 2008, 47, 9200–9211. (40) Nimmanwudipong, T.; Runnebaum, R. C.; Block, D. E.; Gates,
(4) Elliott, D. C. Energy Fuels 2007, 21, 1792–1815. B. C. Catal. Lett. 2011, 141, 779–783.
(5) Furimsky, E. Appl. Catal., A: Gen. 2000, 199, 147–190. (41) Nimmanwudipong, T.; Runnebaum, R. C.; Tay, K.; Block,
(6) Sharma, R. K.; Bakhshi, N. N. Fuel Process. Technol. 1991, 27, D. E.; Gates, B. C. Catal. Lett. 2011, 141, 1072–1078.
113–130. (42) Nimmanwudipong, T.; Runnebaum, R. C.; Block, D. E.; Gates,
(7) Sharma, R. K.; Bakhshi, N. N. Appl. Catal. 1991, 76, 1–17. B. C. Energy Fuels 2011, 25, 3417–3427.
(8) Sharma, R. K.; Bakhshi, N. N. Bioresour. Technol. 1991, 35, (43) Bhore, N. A.; Klein, M. T.; Bischoff, K. B. Ind. Eng. Chem. Res.
57–66. 1990, 29, 313–316.
(9) Sharma, R. K.; Bakhshi, N. N. Energy Fuels 1993, 7, 306–314. (44) Bhore, N. A.; Klein, M. T.; Bischoff, K. B. Chem. Eng. Sci. 1990,
(10) Sharma, R. K.; Bakhshi, N. N. Biomass Bioenergy 1993, 5, 45, 2109–2116.
445–455. (45) Zhao, C.; Kou., Y.; Lemonidou, A. A.; Li, X.; Lercher, J. A.
(11) Vitolo, S.; Seggiani, M.; Frediani, P.; Ambrosini, G.; Politi, L. Angew. Chem., Int. Ed. 2009, 48, 1–5.
Fuel 1999, 78, 1147–1159. (46) S-enol, O. I_.; Ryymin, E. M.; Viljava, T. R.; Krause, A. O. I.
(12) Adjaye, J. D.; Bakhshi, N. N. Fuel Process. Technol. 1995, 45, J Mol. Catal. A: Chem. 2007, 277, 107–112.
161–183. (47) Gutierrez, A.; Kaila, R. K.; Honkela, M. L.; Slioor, R.; Krause,
(13) Adjaye, J. D.; Bakhshi, N. N. Fuel Process. Technol. 1995, 45, A. O. I. Catal. Today 2009, 147, 239–246.
185–202.
(14) Adjaye, J. D.; Katikaneni, S. P. R.; Bakhshi, N. N. Fuel Process.
Technol. 1996, 48, 115–143.
(15) Elliott, D. C.; Hart, T. R.; Neuenschwander, G. G.; Rotness,
L. J.; Zacher, A. H. Environ. Prog. Sustainable Energy 2009, 28, 441–449.
(16) Gagnon, J.; Kaliaguine, S. Ind. Eng. Chem. Res. 1988, 27,
1783–1788.
(17) Samolada, M. C.; Baldauf, W.; Vasalos, I. A. Fuel 1998, 77,
1667–1675.
(18) Gevert, S. B.; Andersson, P. B. W.; Sandqvist, S. P.; Jaeraas,
S. G.; Tokarz, M. T. Energy Fuels 1990, 4, 78–81.
(19) Gevert, B. S.; Otterstedt, J. E. Biomass 1987, 13, 105–115.
(20) Baker, E. G.; Elliott, D. C. ACS Symp. Ser. 1988, 376, 228–40.
(21) Elliott, D. C.; Neuenschwander, G. G. In Developments in
Thermochemical Biomass Conversion; Bridgwater, A. V., Boocock,
D. G. B., Eds.; Blackie Academic & Professional: London, 1997; Vol.
1, pp 611 621.
(22) Piskorz, J.; Majerski, P.; Radlein, D.; Scott, D. S. Energy Fuels
1989, 3, 723–726.
(23) Samolada, M. C.; Baldauf, W.; Vasalos, I. A. Fuel 1998, 77,
1667–1675.
(24) Gevert, B. S.; Otterstedt, J. E.; Massoth, F. E. Appl. Catal. 1987,
31, 119–131.
(25) Petrocelli, F. P.; Klein, M. T. Pet. Sci. Technol. 1987, 5, 25–62.
(26) Laurent, E.; Delmon, B. Appl. Catal., A: Gen. 1994, 109, 77–96.
(27) Centeno, A.; Laurent, E.; Delmon, B. J. Catal. 1995, 154,
288–298.
(28) Centeno, A.; David, O.; Vanbellinghen, C.; Maggi, R.; Delmon,
B. In Developments in Thermochemical Biomass Conversion; Bridgwater,
A. V., Boocock, D. G. B., Eds. Blackie Academic & Professional: London,
1997; Vol. 1, pp 589 601.
(29) Shabtai, J.; Nag, N. K.; Massoth, F. E. J. Catal. 1987, 104,
413–423.
(30) Bredenberg, J. B.; Huuska, M.; Toropainen, P. J. Catal. 1989,
120, 401–408.

4785 dx.doi.org/10.1021/ef2010699 |Energy Fuels 2011, 25, 4776–4785

You might also like